C−F Bond Activation by Aryl Carbocations: Conclusive Intramolecular

Note: In lieu of an abstract, this is the article's first page. Click to increase image size Free .... Ian Mallov , Adam J. Ruddy , Hui Zhu , Stefan G...
0 downloads 0 Views 68KB Size
J. Am. Chem. Soc. 1997, 119, 4319-4320

C-F Bond Activation by Aryl Carbocations: Conclusive Intramolecular Fluoride Shifts between Carbon Atoms in Solution and the First Examples of Intermolecular Fluoride Ion Abstractions

4319

Scheme 1. An Intramolecular Fluoride Ion Shift

Dana Ferraris, Christopher Cox, Rahul Anand, and Thomas Lectka* Department of Chemistry, Johns Hopkins UniVersity 3400 North Charles Street, Baltimore, Maryland 21218 ReceiVed September 3, 1996 ReVised Manuscript ReceiVed January 31, 1997 The chemistry of selective C-F bond activation is undergoing a surge of interest due to the scientific and commercial importance of fluorocarbons.1 However, to date investigations of C-F bond activation in solution usually have involved transition metal complexes or alkali metals. As part of a new approach to C-F bond activation, we were interested in the possibility of using carbocations to abstract fluoride from organic molecules through a three-center transition state or intermediate. Although cationic fluoride transfers involving hypervalent [C-F-C]+ interactions are documented in the gas phase in work by Morton,2 reports of such intermolecular abstractions in solution are virtually nonexistent, and the published cases of intramolecular shifts are fraught with controversy.3 In fact, the question of whether such shifts of fluoride could ever occur in solution has been a source of continuing debate. In this communication, we describe intramolecular fluoride shifts between carbon atoms in solution (Scheme 1) and the first documented examples of analogous intermolecular fluoride abstractions between an aryl cation and its counterion. In the gas phase the reactivity of carbocations is enhanced by the absence of solvent;2 in solution, cations of comparably high reactivity must be generated to provide the driving force for fluoride transfers. It would also be best to generate the carbocations in nonnucleophilic media, employing substrates that contain nonsolvolytically active C-F bonds. Our approach was to use aryl diazonium salts, from which highly reactive aryl cations are known to be generated in solution through mild thermolysis.4 We chose to investigate biphenyl diazonium salts 1 in the belief that upon the formation of aryl cation 2, Ftransfer through a six-membered ring (structure 3) would be favorable (Scheme 1). To minimize trapping of counterion X(1) For a review of C-F bond activation employing metals, see: Kiplinger, J. L.; Richmond, T. G.; Osterberg, C. E. Chem. ReV. 1994, 94, 373. For recent monographs on organic fluorine chemistry see: Synthetic Fluorine Chemistry; Olah, G. A., Chambers, R. D., Prakash, G. K. S., Eds.; Wiley: New York, 1992. Chemistry of Organic Fluorine Compounds II. A Critical ReView; Hudlicky, M., Pavlath, A. E., Eds.; ACS Monograph 187; American Chemical Society: Washington, DC, 1995. (2) (a) Shaler, T. A.; Morton, T. H. J. Am. Chem. Soc. 1989, 111, 6868. (b) Stams, D. A.; Thomas, T. D.; MacLaren, D. C.; Ji, D.; Morton, T. H. J. Am. Chem. Soc. 1990, 112, 1427. (c) Shaler, T. A.; Morton, T. H. J. Am. Chem. Soc. 1994, 116, 9222. (3) In previous fluoride shift claims a dissociative mechanism involving free fluoride in solution could not be ruled out. For an overview see: (a) Olah, G. A. Halonium Ions; Wiley: New York, 1975. (b) Olah, G. A.; Prakash, G. K.; Krishnamurthy, V. V. J. Org. Chem. 1983, 48, 5116. (c) Peterson, P. E.; Bopp, R. J. J. Am. Chem. Soc. 1967, 89, 1283. (d) Ciommer, B.; Scwarz, H. Z. Naturforsch. 1983, 38B, 635. (4) Zollinger, H. Azo and Diazo Chemistry; VCH: Weinheim, 1994; pp 161-277. (5) For a review of weakly coordinating counterions, see: Strauss, S. H. Chem. ReV. 1993, 93, 927. (6) CAUTION! All diazonium salts we report herein are potentially explosive compounds and should be handled with care behind protective shielding both in the solid state and in solution. (7) We measured gas evolution with a Clausen-Kaas microhydrogenationer: Clausen-Kaas, N.; Limborg, F. Acta Chem. Scand. 1947, 1, 884.

S0002-7863(96)03090-9 CCC: $14.00

Table 1. Fluoride-Shifted Products from the Decomposition of Salts 1 yield (%) entry a b c d e

salt

solvent

5

6

7

8

1a 1a 1a 1a 1b

Et2Oa,c Et2Ob,c C6F6b,c C6F6b,d C6F6b,c

20 77

0 0 55 35 49

0 0 0 20 0

0 0 0 0 8

a Reaction run under N gas. b Reaction run under O gas. c Reaction 2 2 quenched by stirring with 10% aqueous NaHCO3. d Reaction mixture purified directly on neutral alumina.

by the aryl cation, we felt a weakly coordinating anion would be necessary.5

Mild heating (40 °C) of diazonium salt 1a6 in ether resulted in N2 evolution over the course of 15 min.7 The reaction was stirred with 10% aqueous NaHCO3 for 15 min, and chromatography over silica afforded ester 5 (20%) as the major product (Table 1, entry a). Decomposition of 1a under an atmosphere of O2 increased the yield of 5 to 77% (entry b). O2 is known to be an effective radical scavenger that shuts down competing chain pathways.8 Heating (40 °C) of 1a in C6F6 instead gave ketone 6 in 55% (entry c). Workup of this reaction by stirring with H218O led to the incorporation of the label into the carbonyl group of 6,9 indicating that hydrolysis is the final step in the fluoride shift process.10 Fluorene trifluoride 7, however, could be isolated along with 6 (C6F6 solvent) upon chromatography of the crude reaction mixture over neutral alumina without aqueous workup in 20% yield (entry d).11 Whereas the ArCF3 moiety is fairly stable to hydrolysis (hot concentrated H2SO4 hydrolyzes ArCF3 over the course of several hours),12 molecules that contain the Ar2CF2 or ArCF2OR moieties are considerably less stable, and can hydrolyze to the corresponding benzophenones and benzoic acids in aqueous acidic solution at ambient temperatures. In our hands, 7 is moisture sensitive and readily hydrolyzes to ketone 6 either in aqueous HCO3- over the course of 0.5 h or on silica gel. Remarkably, when wet 1a (2-3 equiv (8) For a discussion, see: Galli, C. Chem. ReV. 1988, 88, 765. (9) 18O/16O isotope ratios were measured by MS. (10) Free fluoride ion in the aqueous layer after workup was determined by 19F NMR, which revealed roughly the expected amounts (e.g. 2 equiv relative to starting diazonium salt 1a). (11) GC analysis of the crude reaction mixture indicates the generation of 7 in substantial amounts over the course of the reaction. (12) Le Fave, G. M. J. Am. Chem. Soc. 1949, 71, 4148.

© 1997 American Chemical Society

4320 J. Am. Chem. Soc., Vol. 119, No. 18, 1997 of H2O) was employed in the thermolysis,13 ketone 6 was still isolated in 60% yield! Salt 1b, containing the weakly coordinating tetrakis[bis(trifluoromethyl)phenyl]borate (B(ArF)4-) counterion, exhibits chemistry parallel to 1a by affording enhanced yields of fluoride-shifted products under O2.14 A proposed mechanism for the intramolecular reaction is shown in Scheme 1. Loss of N2 from 1 results in formation of aryl cation 215 that abstracts F- from the CF3 group to afford more stable benzylic cation 416 through the [C-F-C]+ transition state or intermediate 3. In ether, 4 is trapped to form an oxonium ion that is hydrolyzed to ester 5 upon aqueous workup.17 Ketone 6 arises from cyclization of cation 4 in C6F6 followed by hydrolysis. Further evidence for an intramolecular fluoride ion shift involving a [C-F-C]+ interaction from the CF3 group to an aryl cationic center is as follows: (1) we observed a secondary kinetic isotope effect (kH/kD) of 1.23 when a 1:1 mixture of 1b and D-labeled isomer 9 reacted to partial completion, consistent with rate-determining formation of an aryl cation;18-20 (2) increased yields of 5 and 6 in the presence of O2 are consistent with the aryl cation mechanism; (3) submitting 2-(trifluoromethyl)biphenyl (10) as an additive to the decomposition of 1 resulted in quantitative recovery of 10; and (4) additionally, MO calculations at the 3-21G level indicate that 4 should be more stable than 2 by 45 kcal/molsa substantial driving force for isomerization.21 In one instance (Table 1, entry e) we observed a small amount of tetrafluoride 8, in which F- must have been abstracted intermolecularly. We speculated that this product may have arisen from B(ArF)4-. Consequently we considered that phenyl cation itself could abstract F- intermolecularly from R3C-F if the C-F bond were present as part of a complex counterion, in which ion pairing could bring the reactive bond into proximity. (13) For example, when salt 1a was made in CH3CN/H218O (3:1), partially dried under vacuum, and decomposed employing the slightly wet solid, the 18O label was incorporated into the carbonyl group of 6 to an 80% extent, which indicates that hydrolysis to ketone 6 occurs primarily under the reaction conditions in this case. (14) Under N2 in C6F6 solvent, 1b gives 20% 5 along with 50% 3,3′,5,5′tetrakis(trifluoromethylbiphenyl), which we interpret to be a radical-derived product from the counterion. (15) Note that the phenyl cation can conceivably react from either the singlet or the triplet state. If the phenyl cation were to react out of the triplet state, it would in all likelihood still abstract F-, not F‚. For example, were the triplet to engage in radical abstraction, it should also react with triplet O2, which the phenyl cation obviously does not, as witnessed by increased yields of fluoride-shifted products in the presence of O2. For a discussion, see: Taylor, R. Electrophilic Aromatic Substitution; Wiley: New York, 1990; pp 215-216. (16) Olah has studied ArCF2+ species under stable ion conditions in solution: Olah, G. A.; Mo, Y. K. J. Org. Chem. 1973, 38, 2686. (17) We monitored the decomposition of 1b in ether by 19F NMR. As the reaction progresses, the CF3 peak of 1b at δ -58.2 (vs CFCl3) diminishes, to be replaced by two other resonances at δ -116.2 (ArF) and δ +65.8 (consistent with the formation of an electron-deficient benzylic substituent ArCF2+ or ArCF2OEt2+, see: Olah, G. A.; Comisarow, M. B. J. Am. Chem. Soc. 1969, 91, 2955). The 19F resonance of B(ArF)4- at δ -63.0 is maintained over the course of the reaction. (18) Large secondary kinetic isotope effects are the hallmark of phenyl cation generation in solution: Swain, C. G.; Sheats, J. E.; Gorenstein, D. G.; Harbison, K. G. J. Am. Chem. Soc. 1975, 97, 791. Swain found that the remote effect due to a deuterium in the 4-position of the diazonium moiety of 1a is negligible. (19) Reaction mixtures were analyzed by 1H NMR and 2H NMR to determine the ratio of labeled to unlabeled reactants and products. (20) Zollinger has found that heterolytic dediazoniations are favored in nonnucleophilic solvents: Szele, I.; Zollinger, H. HelV. Chim. Acta 1978, 61, 1721. (21) We performed MO calculations with the Spartan 4.0 program, Wavefunction Inc.

Communications to the Editor When we heated salt 11 to 60 °C under O2 in C6F6, it gradually discolored (eq 1).22 A vacuum was applied and the volatile

products and solvent were caught in a dry ice trap. Assays of the condensate by GC and NMR showed it to contain PhF (12) (55% yield from 11).23 A secondary kinetic isotope effect kH/ kD of 1.21 for each ortho position was observed24 when a 1:1 mixture of 11 and labeled isomer 13 was pyrolyzed to partial completion in ether solution, again consistent with ratedetermining formation of an aryl cation. Evidence that shifted fluoride originates from a CF3 group of B(ArF)4- was obtained by the isolation of ketone 14, which presumably forms through

alkylation of PhF in the para position,25 protonolysis of the borate counterion by the equivalent of liberated protons,26 and aqueous workup. We believe it is this alkylation that limits the overall yield of PhF. We also found that treatment of NaB(ArF)4 with SbF5 in CH2Cl2 in the presence of PhF produces ketone 14 as the major product, presumably by the same pathway as the reaction of salt 11.27 In conclusion, we have demonstrated C-F bond activation using aryl cations and have provided the first definitive examples of intramolecular and intermolecular fluoride abstractions in solution. Further work is aimed at the intermolecular activation of other fluorocarbons and the synthesis of characterizable fluoronium ions containing hypervalent [C-F-C]+ bonds in solution. Acknowledgment. T.L. thanks the American Cancer Society for a New Faculty Grant, and C.C. thanks JHU for a Marks Fellowship. The authors thank Professors Alex Nickon, Gary Posner, and John Toscano for helpful comments. Supporting Information Available: Spectroscopic details for all new compounds and details on the conduct of decomposition reactions, isotope effect studies, and the synthesis of diazonium salts (10 pages). See any current masthead page for ordering and Internet access instructions. JA963090+ (22) Monitoring of the reaction by 11B NMR showed replacement of the B(ArF)4- counterion at δ -6.60 (vs BF3 in C6F6) by a broad, overlapping resonance at δ -6.40. (23) Boudjouk has shown that silicenium ions can abstract fluoride from the B(ArF)4- counterion: Bahr, S. R.; Boudjouk, P. J. Am. Chem. Soc. 1993, 115, 4514. (24) Once again, we assume a negligible isotope effect for the para position. (25) Fluorobenzene undergoes Friedel-Crafts alkylation primarily at the para position: (a) Friedel Crafts and Related Reactions; Olah, G. A., Ed.; Wiley: New York, 1963. (b) Lichtenberger, J.; Muller, P.; Huguet, M. Bull. Soc. Chim. Fr. 1953, 10, C45. (26) The protonolysis of tetraphenylborate has been investigated: Geske, D. H. J. Phys. Chem. 1959, 63, 1062. (27) See Supporting Information for mechanistic proposals.